Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

download Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

of 90

Transcript of Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    1/90

    TRIBOLOGICAL BEHAVIOR OF NANOCRYSTALLINE METALSPREPARED BY SURFACE MECHANICAL

    ATTRITION TREATMENT

    by

    MICHAEL FRINK

    Presented to the Faculty of the Graduate School of

    The University of Texas at Arlington in Partial Fulfillment

    of the Requirements

    for the degree of

    MASTER OF SCIENCE IN MATERIALS SCIENCE AND ENGINEERING

    THE UNIVERSITY OF TEXAS AT ARLINGTON

    August 2009

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    2/90

    Copyright by Michael Frink 2009

    All Rights Reserved

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    3/90

    ACKNOWLEDGEMENTSI would like to thank my future wife, Donna Chen, for her overwhelming support while I

    was working on my degree. Her encouragement kept me focused and on track to finish my

    degree. I would also like to show my great appreciation to my research advisor, Dr. Efstathios I.

    Meletis, for his advice and guidance throughout my research and studies. I extend my

    appreciation to my committee members Dr. Pranesh Aswath and Dr. Michael Jin for their time,

    effort, and patience. I acknowledge the use of the UTA Characterization Center for Materials

    and Biology (C2MB). Without this facility, analysis of my materials would have been much more

    difficult. I also would like to thank Punnapob Punnakitikashem for helping me to develop the

    thermally evaporated aluminum coating at the UTA Nanotechnology Research & Teaching

    Facility (NanoFAB). Finally, great thanks goes to my fellow group members for all of their help

    and support along the way, especially Dr. Maria Hossou, Cristian Cionea, and Razieh

    Khalifehzadeh.

    May 22, 2009

    iii

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    4/90

    ABSTRACT

    TRIBOLOGICAL BEHAVIOR OF NANOCRYSTALLINE METALS

    PREPARED BY SURFACE MECHANICAL

    ATTRITION TREATMENT

    Michael Frink, M.S.

    The University of Texas at Arlington, 2009

    Supervising Professor: Efstathios I. Meletis

    Nanocrystalline materials have shown great promise for a number of engineering

    applications ranging from structural and electronics to biomaterials and drug delivery.

    Nanoscale grain structures can cause tremendous improvements in strength, hardness and

    other mechanical properties. This is partially due to the suppression of dislocations within

    nanocrystalline grains and the entirely different deformation modes active at that scale. In spite

    of anticipated benefits, very little work has been performed to explore nanoscale effects on the

    tribological behavior of materials. Surface Mechanical Attrition Treatment (SMAT) is an

    effective and economical method to produce nanostructured surface layers in bulk materials.

    The process only takes hours to apply and industry is familiar with the similar process of shot

    peening. In the present work, SMAT was utilized to develop a gradient grain size

    nanocrystalline layer in pure nickel, aluminum, and titanium. The SMAT microstructures

    produced were characterized by surface profilometry, scanning electron microscopy in

    conjunction with energy-dispersive spectroscopy, x-ray diffraction, and nanoindentation. Pin-

    iv

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    5/90

    v

    on-disk wear tests were performed on the SMAT processed materials to study the effect of the

    produced gradient microstructure on the tribological response with comparison to their

    microcrystalline counterparts. Wear testing showed that in general, the nanostructured layer

    reduces the coefficient of friction and improves the wear resistance but the extent of the

    improvements is material dependent. The largest reduction in friction was observed for pure Ni

    (~33%) followed by pure Ti (~14%). Minor effects on friction if any (< 1%) were observed for

    pure Al since its behavior was dominated by aluminum oxide formation at the asperity contacts.

    The SMAT treated Ni and Ti display friction characteristics similar to those of nanocrystalline

    materials possessing a uniform grain size. However, surface roughness resulting from the

    SMAT process may have contributed to somewhat initial higher wear rate. In both of the above

    cases, finer wear debris was produced for the SMAT processed metals suggesting a different

    wear mechanism compared to their microcrystalline counterparts. SMAT aluminum exhibited a

    minor reduction in friction compared to its microcrystalline counterpart and an earlier stage for

    aluminum oxide debris formation during wear. The faster aluminum oxide debris formation

    could be due to surface roughness effects, instead of the applied SMAT layer. The produced

    SMAT layer was not thick enough to have any appreciable effect. The present study shows that

    SMAT processing possesses good potential as a surface nanocrystallization/modification

    method to improve tribological response of certain engineering materials.

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    6/90

    TABLE OF CONTENTS

    ACKNOWLEDGEMENTS ................................................................................................................ iii

    ABSTRACT ..................................................................................................................................... iv

    LIST OF ILLUSTRATIONS.............................................................................................................. ix

    LIST OF TABLES ............................................................................................................................ xii

    Chapter Page

    1. INTRODUCTION............................................ 1

    2. OBJECTIVES ................................................................................................................. 3

    3. REVIEW OF LITERATURE ............................................................................................. 4

    3.1 Nanograin Formation ....................................................................................... 5

    3.1.1 SMAT Surface Mechanical Attrition Treatment ............................. 6

    3.1.2 Physical Vapor Deposition ............................................................... 7

    3.1.3 Electrodeposition .............................................................................. 7

    3.1.4 ECAP Equal Channel Angular Pressing ....................................... 8

    3.2 Wear Mechanisms ........................................................................................... 9

    3.2.1 Fundamental Wear Mechanisms ..................................................... 9

    3.2.1.1 Adhesive Wear ................................................................. 9

    3.2.1.2 Abrasive Wear ................................................................ 10

    3.2.2 Effects from Wear Debris ............................................................... 10

    3.2.3 Aluminum and Titanium Wear Debris Visual Characterization ...... 12

    3.3 Nanograin Formation During Wear ................................................................ 12

    3.3.1 Mechanical Alloying ....................................................................... 14

    3.3.1.1 Tribologically Transformed Structure ............................. 15

    vi

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    7/90

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    8/90

    viii

    5.1 Microstructure................................................................................................. 40

    5.2 Hardness ........................................................................................................ 43

    5.2.1 Nanoindentation ............................................................................. 43

    5.2.2 Knoop Hardness ............................................................................ 47

    5.3 Surface Characteristics .................................................................................. 47

    5.4 Tribological Behavior ...................................................................................... 48

    5.4.1 Pure Nickel ..................................................................................... 50

    5.4.2 Pure Aluminum ............................................................................... 58

    5.4.3 Pure Titanium ................................................................................. 64

    5.5 Tribological Behavior of Nanostructured Metals ............................................ 67

    6. CONCLUSION .............................................................................................................. 69

    REFERENCES ............................................................................................................................... 71

    BIOGRAPHICAL INFORMATION .................................................................................................. 78

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    9/90

    LIST OF ILLUSTRATIONS

    Figure Page

    3.1 Yield stress as a function of grain size [3]. ............................................................................... 5

    3.2 One type of SMAT treatment - Ultra-sonic Shot Peening [8]. .................................................. 6

    3.3 ECAP process: A rod is forced though a 90 bend to refine its structure [14]. ........................ 8

    3.4 Aluminum frictional transients in different relative humidity:(a) 20%, (b) 46%, (c) 65% and (d) 95%. ............................................................................... 11

    3.5 Wear track of cg material undergoes grain refinement during wear [22]. .............................. 13

    3.6 Wear track of nc bulk material undergoes grain coarsening during wear [22]. ...................... 14

    3.7 TTS layer in (a) maraging M250 and (b) Ti15V3Al3Cr3Sn [25]. .................................... 15

    3.8 Comparison of the stress to operate a Frank-Read dislocation source andthe yield stress given by a Hall-Petch relation versus the grain size [5]................................ 16

    3.9 Series of images extracted from a video of an in situ indentation on theultrafine-grained Al film: (a) the indenter is approaching the Al film, (b) bothgrains 1 and 4 grow at the expense of grain 2, (c) grain 2 has been eliminated

    by the growth of its neighbors, and (d) grain 4 continues to grow, resulting ingrain 1 moving back to its original size and grain 4 consuming grain 3on its right side [27]. ............................................................................................................... 18

    3.10 Dimpled fracture surface of Nickel-Tungsten tensile sample with an averagegrain size of 10 nm and with dimple diameters ranging from 100-300 nm [3]. .................... 19

    3.11 Nanovoids forming ahead of a crack in nc nickel. In (d), the inset showsdislocations forming in front of the crack [32]. ...................................................................... 20

    3.12 Tensile test result of nc nickel conducted at various strain rates [3]. ................................... 21

    3.13 Tensile tests performed on various metals at various strain rates [3]. ................................. 21

    3.14 Surface morphologies before and after fretting wear test of polycrystalline nickel;(a, d) 90 m grain size (b, e) 62 nm grain size (c, f) 13 nm grain size [33]. ........................ 23

    3.15 Volume loss versus the inverse Vickers hardness [33]. ....................................................... 23

    ix

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    10/90

    3.16 Micrographs before wear tests: nickel coatings (a-3um, b-250 nm, c-16nm),cobalt coatings (d-2.5 um, e-220nm, f-18nm) and micrographs of samplesafter wear tests in the same order (nickel g-i, cobalt j-l) [34]. .............................................. 24

    3.17 Micrograph of corroded surface of (a) nc and (b) cg copper [41]. ....................................... 26

    3.18 Cyclic fatigue tests of nanocrystalline nickel (20-40 nm grains), ultra-finecrystalline nickel (300 nm grains), and microcrystalline bulk nickel [3]. ............................... 27

    3.19 (a) Vickers hardness of multiple nickel C-2000 samples with varying SMATprocess time. (b) Fatigue life plot for the same samples [45]. ............................................. 28

    3.20 Coefficient of friction of nickel with grain sizes of 8 nm, 22nm, and 61m [4]. ................... 28

    3.21 Variation of fatigue crack growth rate with stress intensity factor in nickel [47]. .................. 29

    3.22 Nickel samples subjected to sinusoidal fatigue loading at a frequency of 10 Hz and a load ratio of 0.3 [48]. ........................................................................................ 30

    4.1 Container used to apply SMAT to samples. ........................................................................... 33

    4.2 Loading profile for nanoindentation test. ................................................................................ 37

    5.1 Optical micrographs of etched (a) Ni-1 (b) Al-2 and (c) Ti-1. ................................................. 40

    5.2 -2 and low angle XRD results for (a) Ni-2 and (b) Al-3. ..................................................... 41

    5.3 -2 scan of Ti-2 before (bulk SMAT Ti-2) and after (SMAT Layer Ti-2) theSMAT surface was removed. ................................................................................................. 42

    5.4 Results of individual indentation tests compared to results of stepped tests

    on SMAT Ni; (a) Reduced modulus and (b) Hardness. ......................................................... 445.5 Nanoindentation hardness results for bulk and SMAT materials. .......................................... 46

    5.6 Reduced modulus results from bulk and SMAT materials. .................................................... 46

    5.7 Typical 3D roughness profiles of materials ............................................................................ 48

    5.8 COF as a function of sliding distance for bulk-Ni and SMAT-Ni. ........................................... 52

    5.9 SEM micrographs showing typical wear track morphology in (a) Ni-1 withhigher magnification of wear track showing brittle nature within inset and(b) Ni-2 [scale bar is 500 m]. ............................................................................................... 53

    5.10 EDS results from nickel wear tracks of Ni-1 and Ni-2. ......................................................... 54

    5.11 COF as a function of sliding distance using a 6.35 mm diameter pin onSMAT nickel compared to testing with a 9.54 mm pin on SMAT and bulk nickel. ............... 54

    5.12 Cross section of S-Ni directly after sputtering [scale bar is 2 m]. ...................................... 55

    x

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    11/90

    5.13 COF as a function of sliding distance for SMAT nickel, N-Ni, andsputtered deposited Ni. ........................................................................................................ 56

    5.14 SEM micrographs showing typical wear track morphology in (a) N-Ni and(b) S-Ni [scale bar is 500 m]. ............................................................................................. 57

    5.15 Wear debris from various nickel samples [scale bar is 10 m]. ........................................... 57

    5.16 Typical 3D wear track profiles of nickel with various microstructures. ................................. 58

    5.17 COF as a function of sliding distance for bulk aluminum (Al-1). .......................................... 59

    5.18 (a) Al-1 Sample after wear tests with track 1 in the center and track 4 on the edge.Pin wear scar morphology for (b) track 1, (c) track 2, (d) track 3, and (e) track 4. .............. 60

    5.19 EDS results from aluminum wear debris in different zones. ................................................ 60

    5.20 COF as a function of sliding distance for bulk and SMAT aluminum. .................................. 61

    5.21 SEM micrographs showing aluminum wear debris from (a) Al-1 track 1,(b) Al-1 track 2, (c) Al-1 track 3, (d) Al-1 track 4, (e) Al-3 track 1, and (f) Al-3 track 2. ........ 63

    5.22 Typical 3D profile of wear tracks in different aluminum microstructures. ............................. 64

    5.23 EDS results from titanium wear debris which displays consistent atomic composition. ...... 65

    5.24 COF as a function of sliding wear of bulk and SMAT titanium. ............................................ 66

    5.25 Typical SEM micrographs showing wear track morphology for (a) Ti-1 and(b) Ti-2 [scale bar is 500 m]. .............................................................................................. 66

    5.26 Typical wear debris appearance in wear track of (a) Ti-1 and (b) Ti-2 ................................ 675.27 Typical 3D view of wear track section from (a) Ti-1 and (b) Ti-2. ........................................ 67

    xi

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    12/90

    xii

    LIST OF TABLES

    Table Page

    4.1 Nominal composition of Nickel-200, wt%. .............................................................................. 31

    4.2 Nominal composition of Aluminum-1100, wt%. ...................................................................... 31

    4.3 Nominal composition of ASTM Grade 2 Titanium, wt%. ........................................................ 31

    4.4 Test sample codes. ................................................................................................................ 32

    4.5 SMAT application times. ........................................................................................................ 33

    4.6 Magnetron sputtering conditions for sputtered nickel and titanium. ....................................... 34

    4.7 Etchants used to reveal grain structure and average grain size of SMAT and bulk Ni, Al, and Ti. ................................................................................................ 35

    5.1 Average grain size of Bulk, electrodeposited, and sputtered samples. ................................. 40

    5.2 Peak location, FWHM, and approximate grain size of SMAT samplesfrom XRD analysis. ................................................................................................................ 42

    5.3 Knoop hardness testing of N-Ni and S-Ni. ............................................................................. 47

    5.4 Surface roughness of materials tested. .................................................................................. 48

    5.5 Hertzian stress of materials studied. ...................................................................................... 49

    5.6 Wear rates of materials tested. .............................................................................................. 50

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    13/90

    CHAPTER 1

    INTRODUCTION

    Wear properties of materials play an important role in any material selection process.

    Due to current material limitations, materials, such as titanium-base and aluminum-base alloys,

    having high strength to weight ratio, cannot be selected for many applications with even a

    moderate tribology requirement due to their poor wear resistance.

    Pure titanium is of special interest in the medical industry due to its biocompatibility.

    However, recent studies have shown that wear debris created by medical titanium alloys

    creates irritation and inflammation around joints [1]. Ideally, pure titanium would be used, but

    pure titanium easily galls and frets when in sliding or vibrating contact with itself or other metals

    and conventional lubricants have trouble overcoming this [2]. Nickel has good tribological

    properties and its alloys are commonly used. However, nickel is expensive, making cost a large

    factor when selecting it for a particular wear application.

    To overcome this, expensive alloys and complex surface coatings are continually being

    developed for use with specific applications. However, little research has been conducted on

    wear improvement through material processing. Processes such as electrodeposition and

    severe plastic deformation can be applied to complex surfaces. Previous research has

    indicated that material wear properties improve as the grain size at the surface drops below 100

    nanometers [3]. Previous studies have shown that FCC nickel, with grain sizes below 100 nm,

    has greater wear resistance than its counterpart with a micrometer grain size [4]. The smaller

    grain size increases hardness and possibly prevents or restricts dislocation motion in the grain.

    This leads to the possible improvements of other FCC metals, such as copper and aluminum,

    and could also indicate an improvement in HCP metals, such as titanium. Surface mechanical

    1

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    14/90

    attrition treatment (SMAT) is a recent surface modification technique used to obtain a

    nanocrystalline microstructure at the surface region of a material. The process is easy to apply

    and produces a gradient grain size nanocrystalline microstructure. This is important because a

    gradient grain size microstructure can produce a gradient in the hardness, resist crack

    nucleation, and resist crack propagation. However, SMAT processing also creates a rough

    surface, which could affect the wear behavior. This study focuses on the tribological effects of

    the SMAT process on pure metals with FCC and HCP structures.

    2

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    15/90

    3

    CHAPTER 2

    OBJECTIVES

    The present study is concerned with the processing-structure-property relationship in

    nanostructured materials. FCC pure nickel, FCC pure aluminum, and HCP pure titanium were

    chosen as model materials to study the wear behavior of a gradient nanocrystalline surface

    structure. More specifically, the objectives are:

    i) study the tribological behavior of gradient nanograined microstructures produced by

    SMAT in three entirely different pure metals;

    ii) to increase the scientific knowledge of the effects of surface processing

    (nanocrystallization and hardening) in regards to tribological behavior.

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    16/90

    CHAPTER 3

    REVIEW OF LITERATURE

    Optimal material properties are continually sought to improve processes and

    applications. Nanocrystalline (nc) materials with grain sizes typically smaller than 100

    nanometers have become the focus of attention due to their unique properties and deformation

    mechanisms. These properties include high strength, strength and/or ductility dependence with

    strain rate, increased resistance to tribological and environmentally-assisted damage, and their

    potential for enhanced super-plastic deformation at lower temperatures with high strain rates [3].

    The properties of nc materials have not been studied in great depth [3], however, there are

    sufficient studies available to provide a general understanding of their potential benefits. Many

    studies compare nc materials to ultra-fine-crystalline (ufc) materials with grain sizes ranging

    from 100 nanometers to 1 micron, and ordinary microcrystalline materials (mc) with grain sizes

    larger than 1 micron.

    One of the benefits of nc materials is described by the Hall-Petch relationship, shown in

    equation 1,

    d

    k yo y

    += (1)

    where y is the yield stress, o is the stress required for dislocation motion (internal friction), ky

    is the strengthening coefficient, and d is the grain size. This relation states that as the grain

    size decreases, the strength of the material increases. This trend is valid for most grain sizes,however, studies have found materials in the lower end of the nc regime seldom follow the Hall-

    Petch relationship [3]. Tests conducted on high quality specimens yield Hall-Petch plots similar

    to Figure 3.1.

    4

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    17/90

    Figure 3.1 Yield stress as a function of grain size [3].It is believed that metallic materials with grain sizes smaller than 10 nm act more

    closely like an amorphous metal, rather than a polycrystalline metal. It is generally accepted

    that the reason for the change in the Hall-Petch slope, observed when grain sizes are below

    100 nm, is suppression of inner grain dislocation mechanisms such as Frank-Read sources [5].

    Other deformation methods, such as grain boundary sliding and nanovoids, are suspected to be

    active in this small domain. Further discussion on these deformation activities will be provided

    later in this chapter.

    In the general nc range, small processing errors can have a great effect on the

    materials properties. Material porosity is a common problem when forming nano-structured

    materials and can be difficult to overcome. These issues are magnified as the desired grain

    size decreases, especially below 10 nm. However, current techniques are beginning to create

    reliable results for materials with grain sizes above 10 nm.

    3.1 Nanograin Formation

    Nanocrystalline materials can be created from three main types of processes; i) from

    isolated nanometer sized particles, ii) from a noncrystalline structure, and iii) from a coarse

    grained (cg) polycrystalline structure [6]. In the first process, nanometer sized particles are

    created, typically by physical vapor deposition (PVD), chemical vapor deposition (CVD),

    5

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    18/90

    electrochemical and hydrothermal methods, and precipitation from a solution, and then

    consolidated to form a bulk material where each particle becomes a nanocrystalline grain. The

    second technique involves starting with an amorphous material and nucleating grains, possibly

    by annealing or mechanical activation [7]. The third process involves increasing a

    polycrystalline materials free energy by introducing numerous material defects and interfaces,

    such as grain boundaries. This can be performed via ball milling, severe plastic deformation,

    sliding wear, irradiation, and spark erosion [8]. Depending on the process used, different grain

    refinement mechanisms transform the polycrystalline structure into a nanocrystalline structure.

    Some common nanocrystalline processes are briefly described below.

    3.1.1 SMAT Surface Mechanical Attrition Treatment

    SMAT involves a grain refinement process with the application of mechanical

    treatments [8]. Repeated localized severe plastic deformation can refine the grain size by the

    creation of shear bands, small-angle grain boundaries, and completely random grain orientation

    [8]. A common SMAT treatment is ultra-sonic shot peening (USSP). This process involves

    shot, typically steel spheres, rapidly hitting the surface of a sample, as shown in Figure 3.2.

    Figure 3.2 One type of SMAT treatment - Ultra-sonic Shot Peening [8]. [

    This process has shown to produce a decrease in grain size in most materials, along

    with an increase in hardness [6, 8-10]. The process is commonly used since its parameters are

    easily controllable, ball-milling equipment can be utilized, and industry is familiar with the

    process [6]. Another benefit of this process is the formation of a gradient microstructure,

    6

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    19/90

    approximately 50 m thick for most materials [11]. Depending on the material, the grains near

    the surface are approximately 10 nm in size, and the grain size gradually increase until the bulk

    material grain size is met. This creates a gradient microstructure and can have beneficial

    applications, especially with tribology and fatigue.

    3.1.2 Physical Vapor Deposition

    Physical vapor deposition is a general name for many thin-film vacuum deposition

    techniques. These processes typically involve plasma sputter bombardment or high

    temperature vacuum evaporation. PVD is commonly used to create thin-films under 100 nm

    thick, but can be used for thicker films. Different PVD processes include evaporative

    deposition, sputter deposition, electron beam physical vapor deposition, cathodic arc deposition,

    and pulsed laser deposition. Evaporative deposition heats the material to be deposited,

    typically through resistive heating, under high vacuum, causing the material to evaporate and

    deposit on the substrate positioned above it. Sputter deposition is performed in a controlled

    ionized atmosphere, typically Argon. The ionized gas particles will be attracted to the positively

    charged target, and upon impact, will knock an atom or cluster of atoms off of the surface. The

    atoms will then deposit on the substrate.

    Due to the complexity of these processes, many processing factors will affect the

    deposited material structure, most notably temperature and pressure [12]. Generally, the

    coatings created with this method are compact, and fully dense. However, porosity and film

    adherence are two common problems.

    3.1.3 Electrodeposition

    Electrodeposition is an inexpensive process where a negative potential is applied to the

    substrate material in an electrolytic bath. This process is easily controllable, and depending on

    process parameters, the grain size of the resulting coating can also be controlled [13]. Complex

    shapes can be electroplated and thicknesses of 100 m or more can be reached. The grain

    size plated can be below 10 nm. Usually, to achieve grain sizes less than 20 nm, a grain refiner

    7

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    20/90

    must be utilized, such as Saccharin [3]. However, this can introduce impurities in the coating.

    Along with this, it is very difficult to produce a fully dense electrodeposited coating. Nanovoids

    can usually be found unless a compaction step is utilized [3]. Compaction will create a slightly

    less dense material coating, which can have a significant effect on the properties of the

    material.

    3.1.4 ECAP Equal Channel Angular Pressing

    Most processes producing nanocrystalline materials create only a thin surface layer.

    However, the ECAP process can produce a bulk nc material. The ECAP process involves

    forcing a cg or single crystal material through a die, typically with a 90 bend ( Figure 3.3). This

    induces large strains in the material leading to grain refinement. A large benefit of the ECAP

    process is the ability to apply a texture to the grains. One pass through an ECAP die will

    produce elongated grains. If the material undergoes many ECAP passes, with rotation before

    each pass, the microstructure can become homogenous [14]. One ECAP pass of a single

    crystal material will produce an ufg polycrystalline material with a preferred texture [15].

    Large Force

    Figure 3.3 ECAP process: A rod is forced though a 90 bend to refine its structure [14].

    8

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    21/90

    3.2 Wear Mechanisms

    Tribology classifies the type of wear that occurs by the relative motion between the

    material surfaces in contact. These motions are typically referred to as sliding, rolling,

    oscillating, impacting, and erosive wear. Tribological tests can also be described with other

    variables based on factors such as the angle of action, lubricants, and environmental

    conditions. Wear tests can also be performed with more than two contacting surfaces, such as

    a ball bearing rolling between two plates and four-ball wear tests.

    The hardness of two materials in contact generally indicates how the softer material will

    wear. A material with a hardness ratio greater than 1.2 with respect to the counterface will tend

    to undergo smooth sliding and less friction. A material with a hardness ratio less than 0.8 with

    respect to the counterface will undergo rough sliding with higher friction coefficients [16].

    3.2.1 Fundamental Wear Mechanisms

    All basic sliding wear mechanisms fall into four categories: abrasive, adhesive, fatigue,

    and oxidative wear [17]. Commonly, a wear system will not be a single type of wear, but a

    combination of mechanisms. Adhesive and abrasive wear are the two dominant wear

    mechanisms in this study, and are briefly described below.

    3.2.1.1 Adhesive Wear

    Other names are sometimes used to describe adhesive wear, such as scoring, galling,

    and seizing, but they all describe the same process. As one surface slides relative to another,

    surface asperities from the roughness of both materials contact, and create very high local

    pressures. This localized high pressure creates plastic deformation and possibly welding

    between the two surfaces. As the sliding continues, the welded region will typically break from

    the softer material, creating a void on the surface and a protrusion on the second surface.

    Debris is ultimately created when the welded asperities break, and adds to the wear of the

    materials. The debris formation typically leads to abrasive wear.

    9

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    22/90

    Archards law can be used to describe the wear volume per unit sliding distance, W

    [18],

    H KPW = (2)

    where P is the applied load, H is the hardness of the softer material, and K is the wear

    coefficient. K is a relative comparison between the relative strength of the interface junction and

    the strength of the asperities that make up the junction. A low value of K indicates that the

    created junction material is softer than the surface asperities on both materials. Many factors

    affect the wear coefficient, such as material hardness, roughness, and material type. Typically,

    similar materials and very soft materials are the most susceptible to adhesive wear due to acold welding mechanism. Lubrication can greatly decrease the amount of adhesive wear

    because of the thin film created that separates the two surfaces.

    3.2.1.2 Abrasive Wear

    Abrasive wear occurs when hard particles remove material from one or both of the

    contacting surfaces. The hard particles can be embedded in one of the contacting surfaces,

    such as in diamond saw blades, and is called two-body abrasive wear. The hard particles can

    also be loose, between the two contacting surfaces. This process is called three-body abrasive

    wear. The friction properties can differ greatly based on whether two-body wear or three-body

    wear is present. In two-body wear, lubrication can actually increase the wear rate because it

    squeezes out dull particles in the second material, and exposes a fresh sharp particle.

    3.2.2 Effects from Wear Debris

    Wear debris can have a strong effect on the wear process by turning a two-body wear

    process into a three-body wear process. This will affect the materials lifetime. Studying the

    composition and structure of the wear debris allows for a better understanding of what is

    occurring during the wear process.

    An excellent study example is aluminum. The tribological properties greatly depend on

    the alloy content and the environment [19]. However, the mechanism of debris formation is the

    10

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    23/90

    same. Many researchers in the past assumed that wear debris created from aluminum

    substrates is aluminum oxide [19]. This is assumed because the current wear theory predicts

    that aluminum will be worn from the surface and oxidize due to the localized high temperatures.

    However, this theory does not explain observations seen in reference [19]. Figure 3.4 shows

    that there is a reduction in the friction coefficient followed by a rise before it reaches steady

    state.

    Figure 3.4 Aluminum frictional transients in different relative humidity: (a) 20%, (b) 46%, (c) 65%and (d) 95%.

    The amount of humidity present during the tests affects how soon the dip is seen. As

    the humidity increases, the dip occurs faster. This behavior cannot be simply attributed to the

    formation of aluminum oxide but to the formation of both aluminum oxide and aluminum

    hydroxide [19]. When a critical thickness of aluminum oxide is reached, a sudden drop in thefriction occurs, and wear debris begins to form. As the wear debris mechanically mixes with

    moisture from the atmosphere, aluminum hydroxide is produced, and a rise in friction occurs.

    This study indicates that mechanical mixing does not only occur between the two contacting

    11

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    24/90

    surfaces, but also with the atmosphere. This phenomenon can also be seen in steel with iron

    hydroxides forming in high relative humidity environments, acting as a lubricant to reduce the

    wear rate and decrease the friction coefficient [20].

    3.2.3 Aluminum and Titanium Wear Debris Visual Characterization

    Analyzing the wear debris from aluminum and titanium can be difficult because their

    debris is commonly in an amorphous or nanocrystalline form. As was shown in the previous

    section, aluminum wear debris forms aluminum oxide and then after mechanical mixing with

    moisture from the atmosphere, forms amorphous aluminum hydroxide. Before careful analysis

    is conducted, the debris can be preliminarily visually characterized by its color. Aluminum oxide

    will appear grey to silver whereas aluminum hydroxides will appear black.

    Titanium wear debris is typically titanium dioxide. Titanium dioxide exists in three

    crystalline polymorph phases at room temperature; rutile (tetragonal crystal structure), anatase

    (tetragonal crystal structure), and brookite (orthorhombic crystal structure). When pure titanium

    oxidizes in humid air, the formation of rutile is thermodynamically favored [21]. Rutile appears

    as a dark red to black color, anatase is an indigo to black color, and brookite is typically an

    orange color. This basic knowledge can greatly help in the identification of the debris,

    especially when only small quantities are created.

    3.3 Nanograin Formation During Wear

    When a cg material undergoes sliding wear, the large strain present at asperity contact

    points will create a large amount of dislocations. As plastic strain grows in the bulk grains,

    dislocations build up in the grain walls and the grain starts to shrink [16]. At the highest strains,

    the grain structure will be refined with sharp grain boundary walls which resemble grains

    produced at high temperature. These nano-sized grains will increase with depth until the bulk

    grain size is obtained. As the sliding contact continues, the layer becomes thicker, but the

    exponential decay remains the same [16]. Dislocations start to form at the interface of the

    nano-structure surface layer. Cracks begin to grow at this interface, and as the cracks reach

    12

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    25/90

    the surface, large portions of material is removed creating large debris [22]. As debris formation

    increases, the wear rate will also increase. Figure 3.5 shows a cross-section of a typical

    steady-state wear track. The bulk material has coarse grains, whereas, the wear track region is

    made up of grains in the nano-scale. In the transition zone between the nano-scaled grains and

    the bulk coarse grained material, a large dislocation density exists.

    Dislocation wallsnc surface grains

    Coarse rains

    Figure 3.5 Wear track of cg material undergoes grain refinement during wear [22].

    The transferred material created during adhesive wear can be further deformed and

    mechanically mixed with counterface material and/or environmental components. This results

    with the transferred material possessing a very fine or nanostructured grain size [23].

    Depending on the material, this can also lead to a thin-film layer forming on the wear region,

    changing its wear characteristics. The thickness of this layer, typically called the tribologically

    transformed structure (TTS), can be dependent on the size of the initial wear debris created,

    ultimately affecting the wear characteristics. Mechanical alloying and TTS formation is

    commonly used to describe the wear track and wear debris on coarse grained materials, and is

    described in the following sub-sections.

    If the grains at the surface are in the nano-scale regime before wear occurs,

    dislocations will become difficult to form and less dislocation cracks will form at the surface.

    13

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    26/90

    The reduction in dislocations, especially near the surface, will prohibit nano-grains from

    becoming loose debris. Friction at the surface will cause some grain growth, primarily by grain

    boundary sliding and grain rotation, but also by temperature induced annealing. Figure 3.6

    displays the resulting wear track morphology. This growth mechanism prevents debris

    formation by suppressing dislocations, allowing for a decrease in the wear rate of the material.

    It is the aim of this report to better understand the formation of this layer and its effects on

    tribological properties.

    Figure 3.6 Wear track of nc bulk material undergoes grain coarsening during wear [22].

    3.3.1 Mechanical Alloying

    A growing theory of Mechanical Alloying [24] has been used to explain how a

    nanocrystalline surface layer is created on a substrate. Mechanical alloying is a process where

    two materials are mixed together by a combination of high strain and locally high temperatures.

    The alloy will have a large volume fraction of dislocations, resulting in the nucleation of nano-

    sized grains in the debris or as a surface layer. This process can occur in the two-body or

    three-body wear regimes. As a material is worn, material transfer can occur with little debris

    formation, by diffusion or adhesion, creating a mixed layer at the surface of both counter

    materials. In three-body wear, debris is formed, mixed with high strain rates, creating a new

    third phase. Mechanical Alloying has important consequences in the wear of a material. The

    14

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    27/90

    surface layer, or debris, created can act as a lubricant or an abrasive, drastically changing the

    tribological properties of the materials.

    3.3.1.1 Tribologically Transformed Structure

    As a material undergoes initial wear, in particular fretting wear (small oscillations,

    usually smaller than the contact radius), a deformed region can form on the surface. This

    region forms due to high strains between the two contacting surfaces and subsequent

    recrystallization of the surface region. This new phase formed is the Tribologically Transformed

    Structure (TTS) layer [25]. Figure 3.7 displays a TTS region in a maraging steel and a titanium

    alloy. The sliding direction is horizontal. InFigure 3.7b, cracks can be seen running from the

    TTS layer through the bulk material along the materials slip plane.

    TTS layer

    Figure 3.7 TTS layer in (a) maraging M250 and (b) Ti15V3Al3Cr3Sn [25].

    The TTS layer starts to form as a thin layer on the substrate near the initial wear cycles.

    As the wear cycles increase, the thickness increases until it reaches a steady state value.

    Debris formation is minimal during this formation stage, however, once the critical thickness is

    reached, debris will be created at the same formation rate as the TTS layer. The maximum

    thickness of this layer will depend on the substrate and pin microstructures and the testingparameters. Due to high strains between the contacting surfaces, the TTS layer will have a

    nanocrystalline microstructure, along with the debris that is created.

    15

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    28/90

    3.3.2 Deformation Mechanisms

    To better understand wear properties of nc metallic materials, some of the proposed

    deformation mechanisms need to be understood. Metallic materials with grains smaller than

    100 nm suppress dislocation motion within the grains. As the grain size decreases, the stress

    required to activate dislocation motion is higher than alternate deformation methods. Figure 3.8

    displays the stress required to activate Frank-Read sources in copper and the flow stress as

    calculated from the Hall-Petch relation. This trend suggests that dislocations emitted from the

    grain are not the dominating deformation mode. Three proposed methods to explain the

    deformation behavior include Grain Boundary Sliding, Grain Rotation, and Nanovoids.

    Figure 3.8 Comparison of the stress to operate a Frank-Read dislocation source and the yieldstress given by a Hall-Petch relation versus the grain size [5].

    3.3.2.1 Grain Boundary Sliding and Rotation

    Grain boundary sliding and grain boundary rotation are usually tied to one another.

    These deformation methods consist of the formation and annihilation of dislocations in the grain

    boundaries. This process causes grains to grow and rotate as they deform.

    To better understand the process, Ke et. al. [26] strained two samples of silver, one with

    10 nm grains and the other with 110 nm grains. The results show no evidence of dislocations in

    16

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    29/90

    the sample with 10 nm grains and clear evidence of dislocation motion in the 110 nm grain

    sample. The deformation in the 10 nm grain sample is believed to be caused by dislocation

    motion within the grain boundaries.

    The grain growth mechanism can be explained using thermodynamic principles. The

    free energy of the system can be lowered by the growth of grains favorably oriented to respond

    to the indentation load [27]. The grain reconfigurations are made possible, kinetically, by the

    fact that the grain boundaries are very mobile even at room temperature [27]. This mobility may

    result from the absorption of dislocations in the boundary, boundary sliding, or the simple

    application of stress to the boundary [27]. This has been captured using TEM while deforming

    an ufc aluminum film (Figure 3.9). Further tests performed by Jin et. al. [27] on iron did not

    display the grain growth observed in aluminum. Nano-sized precipitates in the grain boundaries

    pinned the grains preventing any sliding or rotating. Therefore, the composition and

    microstructure of the nc metallic materials must be considered.

    A further study by Vevecka et. al. [28] on the cyclic behavior of lead found that grain

    boundary sliding occurred in low cycle fatigue, causing cracks to form around the grain

    boundaries. These results were consistent with other results performed on lead and other

    metallic materials.

    Simulations conducted in the nc regime showed that the stress-strain response of the

    nc metallic material exhibited an inelastic response at the grain boundaries, and an elastic

    response in the interior of the grains [29]. Plastic deformation was rarely evident inside the

    grains before failure. Instead, cracks would start to form as the grains reoriented, increasing the

    internal strain until failure. The grain boundary cracking accounts for the decrease in

    macroscopic ductility as a metallic materials grain size decreases.

    17

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    30/90

    Figure 3.9 Series of images extracted from a video of an in situ indentation on the ultrafine-grained Al film: (a) the indenter is approaching the Al film, (b) both grains 1 and 4 grow at the

    expense of grain 2, (c) grain 2 has been eliminated by the growth of its neighbors, and (d) grain4 continues to grow, resulting in grain 1 moving back to its original size and grain 4 consuming

    grain 3 on its right side [27].

    3.3.2.2 Nanovoids

    In nc metallic materials, the fracture surface of fatigued specimens contains spherical

    dimples, usually a magnitude larger in diameter than the average grain size. The presence of

    large dimples suggests that there is considerable local plastic deformation in the material as a

    crack is progressing. Figure 3.10 shows a sample of nickel-tungsten with an average grain size

    of 10 nm and spherical dimples of 100 to 300 nm in diameter [3].

    18

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    31/90

    Figure 3.10 Dimpled fracture surface of Nickel-Tungsten tensile sample with an average grainsize of 10 nm and with dimple diameters ranging from 100-300 nm [3].

    [Tests have been performed to explain the dimpling of the surface. As a crack is

    propagating through a nc metallic material, dislocations have been seen at the crack tip [3, 30-

    32]. The dislocations emitted from the tip act to blunt the crack. They travel from the crack tip

    to the grain boundaries. As the elastic energy grows in the grain boundaries, the triple junctions

    will experience a high stress state. Grain boundary sliding can occur with the dislocation motion

    within the boundary, and a small void will form at a triple junction. As the crack advances

    closer, the void will grow until the crack reaches it. Nanovoids do not form at every triple

    junction in front of a crack. They will only grow when the elastic energy at the triple junction

    reaches a stress limit. This process can be seen in Figure 3.11, in a nc nickel. Figure 3.11(a)

    shows three voids at three triple junctions in front of the crack tip. The crack grows into these

    nanovoids. Figure 3.11(d) shows evidence of dislocation motion in front of the crack tip,

    reinforcing previous simulations.

    19

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    32/90

    Figure 3.11 Nanovoids forming ahead of a crack in nc nickel. In (d), the inset showsdislocations forming in front of the crack [32].

    3.4 Tribological Behavior of Nanostructured Metals

    3.4.1 Material Behavior

    As previously stated, the nc metallic materials display unique properties. Strain rate

    dependent deformation is usually only observed in polymers and not in metals. However,

    Figure 3.12 displays strain rate dependence of nc nickel with an average grain size of 40 nm.

    This trend is not only seen in nc nickel, but in most nc metallic materials. Figure 3.13

    displays results from a variety of tests performed on mc (black data points), ufc (blue data

    points), and nc (red data points) metallic materials. Figure 3.13 shows very little change in yield

    strength at low strain rates for mc and ufc metallic materials but shows a strain rate dependence

    for the nc materials. However, at very high strain rates above 1000 s -1, there is a dependence

    20

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    33/90

    on strain rate for all materials. When a load is applied extremely fast to any material, it will

    prevent dislocation motion and delay failure. Nanocrystalline materials already suppress

    dislocations due to their small grain size, and therefore display a strain rate dependence. This

    strain rate dependence could have various implications to structural applications, where low

    strain rates, such as below 0.1 s -1, are observed.

    Figure 3.12 Tensile test result of nc nickel conducted at various strain rates [3].

    Figure 3.13 Tensile tests performed on various metals at various strain rates [3].

    21

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    34/90

    3.4.1.1 Wear Resistance

    Sliding wear in mc metallic materials creates large plastic strains in the grains at the

    surface [24]. Strain at the surface of the material causes grain refinement with dislocation

    mechanisms, creating a rough surface and wear debris. This leads to cracks forming at the

    surface and reducing the metallic materials life. Since inner granular dislocation motion is

    suppressed with nc metallic materials, the wear rate and crack formation decrease as compared

    to mc metallic materials. Results from wear tests performed by Jeong et. al. [33] on mc nickel

    and electroplated nc nickel are shown in Figure 3.14.

    It must be noted that the surface topography is not consistent between the 90 m nickel

    (Figure 3.14 (a, d)) and the 13 nm nickel ( Figure 3.14 (c, f)), however, the fretting observed

    along the wear track is a good indication of the wear resistance. Comparing (d), (e), and (f),

    deep gouges along with fretting can be seen in the mc sample (a, d). As the grain size

    becomes smaller, as in (f), the wear track becomes shallow and minimal fretting is observed.

    This correlates well with the Hall-Petch relationship: as the grain size decreases, the strength

    and hardness increases. The decrease in the gouging effect will also have a positive impact on

    the fatigue life of the material by preventing crack nucleation sites. This test also displayed a

    decrease in the friction coefficient from the mc nickel to the nc nickel by over 50%. The

    increase in wear resistance can also be seen in Figure 3.15, where the volume loss of each

    sample was plotted against the Vickers hardness.

    22

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    35/90

    Figure 3.14 Surface morphologies before and after fretting wear test of polycrystalline nickel; (a,d) 90 m grain size (b, e) 62 nm grain size (c, f) 13 nm grain size [33].

    Figure 3.15 Volume loss versus the inverse Vickers hardness [33].

    A similar test was performed by Wang et. al. [34]. A 50 m thick nickel and cobalt layer

    was electrodeposited on a substrate. Figure 3.16 displays the electroplated nickel before pin-

    on-disk wear tests (a-c) and after the wear tests (g-i) as well as the electroplated cobalt before

    wear tests (d-f) and after wear tests (j-l). As in Figure 3.14, the wear resistance of the nickel

    23

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    36/90

    samples increases as the grain size decreases. The material becomes harder and the

    mechanisms of deformation are suppressed as the grain size decreases, resulting in the

    decrease in wear rate. The cobalt samples also display a slight decrease in wear rate with

    decreasing grain size, but at a much smaller scale. The wear rate for both the mc and nc

    cobalt coatings are far lower than the nickel coatings. This can be attributed to better resistance

    of HCP structure to adhesion interactions compared to other atomic structures [34].

    (g)

    (h)

    (i)

    (j)

    (k)

    (l)

    Figure 3.16 Micrographs before wear tests: nickel coatings (a-3um, b-250 nm, c-16nm), cobaltcoatings (d-2.5 um, e-220nm, f-18nm) and micrographs of samples after wear tests in the same

    order (nickel g-i, cobalt j-l) [34].

    Only a handful of studies have explored the grain growth phenomenon in

    nanocrystalline materials subjected to wear. Nanocrystalline aluminum has shown a decrease

    in wear debris compared to mc aluminum at high loads and high sliding velocities [35].

    However, large plastic strains exceeding 10 (1000%) are known to be present at the interface[16], and the decrease in wear debris can partially be attributed to its increased shear strength

    with decreasing grain size.

    24

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    37/90

    Ultra-fine grained titanium, created using an ECAP technique, showed improved friction

    and lower wear rates compared to an annealed coarse grained titanium. Titanium oxide was

    produced on both the cg and ufg titanium specimens. It is also important to note that the wear

    mechanisms tended to form micro-plowing and delamination of the surface layer. This suggests

    that high contact stresses are present just below the contact surface, and fatigue affected both

    ufg and cg titanium. More titanium oxide was found on the cg wear track when compared to the

    ufg wear track. Even though their friction coefficients were similar, the ufg had a lower wear

    rate [36].

    3.4.1.2 Corrosion Resistance

    A possible added benefit of a nanocrystalline surface layer is an improvement in a

    materials corrosion resistance. A consensus has not been made on whether the properties

    improve, degrade, or lie somewhere between. The corrosion resistance of electroplated nickel

    has been found to improve with deceasing grain size [37]. This is due to the ability of the

    surface to rapidly form a passivating layer of nickel hydroxide since the surface energy is high

    from the exposed grain boundaries.

    In titanium, Balyanov et. al. [38] concluded that ufg titanium (approximately 300 nm)

    improved the corrosion resistance by rapidly creating a passivating layer, due to the higher

    surface energy. However, Garbacz et. al. [39] showed that nc titanium had little to no effect on

    the corrosion resistance, and actually degraded the corrosion resistance. [40][41]

    In copper, generally it has been found that the corrosion resistance decreases with nc-

    copper [40,41]. However, voids forming during the processing of the nc copper were suspected

    as one reason for the decrease in the corrosion resistance [41].

    Finally, Zhang et. al. [42] concluded that nc zirconium exhibited better corrosion

    resistance compared to mc zirconium, Mordyuk et. al. [43] reported that ultrasonic shot peened

    austenitic steel showed improved corrosion resistance by increasing the volume fraction of

    25

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    38/90

    strain-induced martensite, and Aghion et. al. [44] concluded that the corrosion resistance of nc

    magnesium is lower than that of the mc magnesium.

    All of the aforementioned reports agree on two points: the large area fraction of grain

    boundaries present at the surface of nc-materials allow the material to corrode more uniformly

    and pitting is more prevalent in cg-materials since impurities tend to segregate to grain

    boundaries. Figure 3.17 shows the surface morphology of nc copper and cg copper. The nc

    copper had a slightly higher corrosion rate, however, as can be seen in the micrograph, the

    surface corroded much more uniformly.

    Figure 3.17 Micrograph of corroded surface of (a) nc and (b) cg copper [41].

    3.4.2 Fatigue Response

    Nanocrystalline metallic materials show also an improvement in the cyclic response.

    Fatigue tests shown in Figure 3.18 were conducted at a R-value of zero and a frequency of 1

    Hz, where R is the applied stress ratio of minimum to maximum stress. The results show an

    increase in the high cycle fatigue strength of nc nickel. However, further experiments have also

    shown that metallic ufc materials undergo cyclic softening and a weakening of their low-cyclefatigue resistance [3]. Coarse grained metallic materials tend to resist low cycle fatigue better

    than nc grained materials but are less resistant to high-cycle fatigue.

    26

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    39/90

    Figure 3.18 Cyclic fatigue tests of nanocrystalline nickel (20-40 nm grains), ultra-fine crystalline

    nickel (300 nm grains), and microcrystalline bulk nickel [3].Many studies have shown that SMAT introduces plastic deformation in the surface of a

    material creating a nc zone and producing compressive stresses near the surface. A grain size

    distribution will be created where nc grains will be developed at the surface followed by a grain

    size increase with distance from the surface [45]. This can increase the life of the material by

    suppressing the tensile zone in front of any stress riser, preventing cracks from forming.

    However, it must be noted that applying a SMAT process for too long will provide no benefit and

    could degrade the properties of the material as seen in Figure 3.19 (b). The SMAT application

    applied for 30 minutes has a much higher fatigue limit than the original or other SMAT

    processes. Figure 3.19a shows that there is little difference in hardness values for the varied

    times. The properties will degrade because dislocations, microcracks, and surface

    contaminates will increase as the process time increases creating nucleation points for cracks.

    In stainless steels, annealing can be conducted after the SMAT process to relax the

    compressive stress while maintaining the nc zone at the surface. This will lead to improved

    fatigue life through a combination of strain hardening and strain-induced martensitic

    transformation [46].

    27

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    40/90

    (a) (b)

    Figure 3.19 (a) Vickers hardness of multiple nickel C-2000 samples with varying SMAT processtime. (b) Fatigue life plot for the same samples [45].

    Figure 3.20 displays the improved coefficient of friction (COF) for nickel as the grain

    size decreases. The lower COF produces less wear and hence less gouging will be present,

    eliminating crack nucleation sites.

    Once a crack forms, it will inevitably grow. The crack growth rate in nc metallic

    materials is larger than the growth rate of mc metallic materials ( Figure 3.21). In this graph, the

    nc material shifts the curve to the left, similar to increasing the load ratio (R). The load ratio is

    the minimum stress divided by the maximum stress. As the ratio is increased, a larger is

    applied, and the crack propagation rate will increase.

    Figure 3.20 Coefficient of friction of nickel with grain sizes of 8 nm, 22nm, and 61m [4].

    28

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    41/90

    Figure 3.21 Variation of fatigue crack growth rate with stress intensity factor in nickel [47].

    Multiple factors exist to explain the increase in crack growth. First, as stated previously,

    grain boundary rotation, sliding, and nanovoids become the dominate deformation process in a

    nc metallic material. This allows nanovoids to form ahead of the crack, creating a path for the

    crack to follow. Second, with small grains, the active slip planes in each nano-sized grain will

    only slightly deflect the crack. This enables a crack to propagate in a relative straight path,

    mimicking a brittle fracture. This reduces any crack closure mechanisms that are present in cg

    materials. Figure 3.22 compares cracks in mc, ufc, and nc nickel. As a crack propagates along

    large granular slip planes, it will create a saw tooth type path, as seen in Figure 3.22a. When a

    material is stressed, the crack faces can come in contact with each other, reducing the stress at

    the crack tip, and slowing the crack growth rate. However, with nc metallic materials, the grains

    are small enough to keep the crack deflection minimal, keeping the crack strictly in mode I

    fracture. Simulations of crack growth in nanocrystalline metallic materials have also been

    conducted [48]. The models predict that as the grain size becomes smaller, the rate of crack

    growth will increase.

    29

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    42/90

    Figure 3.22 Nickel samples subjected to sinusoidal fatigue loading at a frequency of 10 Hz and

    a load ratio of 0.3 [48].

    Since the majority of fatigue cracks initiate at the surface, an ideal material will have a

    nc surface layer with a gradient microstructure, with the grain size increasing with the distance

    from the surface, similar to the microstructure created by the SMAT process.

    30

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    43/90

    CHAPTER 4

    EXPERIMENTAL

    4.1 Materials

    The selected materials used were nickel-200, Aluminum-1100, and ASTM (American

    Society for Testing and Materials) Grade 2 Titanium. All bulk model metals were selected

    because their purity is greater than 99% and they are commonly used in tribological studies.

    The material purity allows for accurate observation of the tribological behavior of their single

    element form. The nominal composition of the bulk materials is shown in the following tables

    [49].

    Table 4.1 Nominal composition of Nickel-200, wt%.

    Ni C Mn Fe Si Cu99.5 0.06 0.25 0.15 0.05 0.05

    Table 4.2 Nominal composition of Aluminum-1100, wt%.

    Al Si plus Fe Cu Mn Zn Other 99.0 min 0.95 0.05 to 0.20 0.05 0.10 0.15

    Table 4.3 Nominal composition of ASTM Grade 2 Titanium, wt%.

    Ti C H O N Fe99.3 min 0.10 0.0125 0.25 0.03 0.30

    The bulk Nickel-200 and Aluminum-1100 was ordered in the form of 38.1 mm diameter

    bar stock. The Grade 2 Titanium was ordered in the form of 51.5 cm diameter bar stock. Discs

    38 mm in diameter disks and 1.6 mm thick were cut on a lathe for preparing bulk and SMAT

    samples.

    31

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    44/90

    A large number of samples have been created and labeled for easier identification.

    Table 4.4 displays the sample name, material, and the way it was made and/or treated. The

    following sections will detail the way the samples were prepared and treated.

    Table 4.4 Test sample codes.

    Code Material DescriptionNi-1 Nickel Coarse grain disk from bar-stockNi-2 Nickel SMATN-Ni Nickel ElectrodepositedS-Ni Nickel Magnetron Sputtered Al-1 Aluminum Coarse grain disk from bar-stock Al-2 Aluminum Coarse grain disk from bar-stock Al-3 Aluminum SMAT

    T-Al Aluminum Thermally evaporatedTi-1 Titanium Coarse grain disk from bar-stockTi-2 Titanium SMATS-Ti Titanium Magnetron Sputtered

    4.1.1 Bulk Specimens

    The nickel, aluminum, and titanium bulk samples were ground with silicon carbide

    papers down to 800 grit to remove any machining marks and layers with residual stresses.

    They were then progressively polished down to 0.25 m diamond paste with methanol. This

    was the final polish step for aluminum. Polishing with 0.05 m alumina paste and methanol wasthe final step for nickel and titanium samples. All samples were cleaned with methanol in an

    ultrasonic cleaner for 15 minutes after their final polish.

    4.1.2 SMAT Specimens

    The SMAT samples were polished and cleaned following the same steps as the bulk

    samples. To apply SMAT to the sample discs, a Spex 8000M Mixer/Mill was utilized. A Spex

    8001 hardened steel vial, typically used to ball mill powders, was modified to hold the sample

    discs. Four 3.18 mm diameter 440 steel balls were used for the treatment at a frequency of 18

    Hz.

    32

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    45/90

    Cap

    Backing Plate

    Retention Ring

    Sample

    Container

    Ball

    Figure 4.1 Container used to apply SMAT to samples.[

    The treatment times differed for the different materials. Many SMAT treatment times

    can be found in the literature for aluminum [9, 22, 50, 51] and titanium [10,11] and the SMAT

    application time for the present work is based on those findings. However, very little research

    has studied the optimal SMAT application times for nickel. Therefore, the nickel SMAT

    application time was based on work performed by Chun Sheng-Wen et. al. [52], and can be

    verified by a recent publication studying different SMAT application times on nickel [53]. The

    SMAT times used in the present work are shown in Table 4.5 .

    Table 4.5 SMAT application times.

    Sample Code Time (min)Nickel-200 Ni-2 60

    Aluminum-1100 Al-3 15Titanium Grade 2 Ti-2 60

    4.1.3 Electrodeposited Nickel

    An electrodeposited nickel sample was also purchased for comparison purposes. The

    electrodeposition conditions are described elsewhere [54]. The sample was deposited with a

    current density of 15 mA/cm2 at a temperature of 30C. Saccharin, Coumarin, and Lauryl

    sulfate were used in the bath composition to help create a more dense material. The resulting

    33

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    46/90

    material is a fully dense nanocrystalline nickel sheet 1.5 mm thick. The sample is labeled N-Ni.

    It was polished similarly as the bulk and SMAT nickel.

    4.1.4 Thin-film Samples

    To help evaluate the benefits of a nanocrystalline surface layer in tribological

    applications, thick-filmed samples were also prepared. The nickel and titanium samples were

    prepared using a magnetron sputtering technique, and the aluminum sample was prepared by

    thermal evaporation. Each sample is approximately 1 m thick.

    4.1.4.1 Magnetron Sputtering

    Following well-known and established techniques, particularly developed by Thorton

    [12], multiple nickel and titanium samples were prepared to determine the optimal coating

    parameters using the magnetron sputtering equipment in the Surface and Nano Engineering

    Laboratory. Silicon wafers with a {111} orientation were used as a substrate. Each substrate

    was RF sputter cleaned by applying a 35 W bias to the substrate for 15 minutes in an argon

    atmosphere of 6 mtorr. This removed any oxide layer from the substrate surface to provide for

    a better bonding surface. Also, each target was cleaned by pre-sputtering while the shutter

    remained above the target, covering the sample. The sputtering deposition for both samples

    was conducted with 60 W of DC power applied to the target and 40 W of RF power applied to

    the substrate. The nickel was sputtered directly on silicon substrate, whereas the titanium was

    sputtered on a thin layer of chromium, which acted as a bond layer. The deposition conditions

    for the latter two polycrystalline samples are shown below in Table 4.6 .

    Table 4.6 Magnetron sputtering conditions for sputtered nickel and titanium.

    SampleBase

    Pressure

    (torr)

    SubstrateTemperature

    (C)

    Rotation

    (RPM)

    Time

    (min)

    Thickness

    (m)S-Ni 5.9x10-7 400 10 165 1.00S-Ti 4.8x10-7 550 10 360 0.90

    34

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    47/90

    4.1.4.2 Thermal Evaporation

    Magnetron sputtering of aluminum thin films can result in a porous and columnar

    structured film. Therefore, thermal evaporation was utilized. Following Semaltianos [55], a high

    deposition rate was used to create a thick aluminum coating on a silicon wafer substrate. A

    Varian NRC 3117 thermal evaporator was used and the aluminum was deposited at the

    evaporators maximum deposition rate of 4 /s and a pressure of 10 -5 torr.

    4.2 Microstructure

    4.2.1 Grain Size

    The grain size of the bulk material was determined by following the Abrams Three-

    Circle Procedure described in ASTM E112-96. Table 4.7 displays the etchant used on each

    material.

    Table 4.7 Etchants used to reveal grain structure and average grain size of SMAT and bulk Ni, Al, and Ti.

    Material ASTM E407-07Etchant Composition

    Nickel 13350 mL HNO3 50 mL Acetic

    Acid

    Aluminum 2 3 mL HF100 mL water

    Titanium 18710 mL HF30 mL HNO3 50 mL water

    The approximate grain size of N-Ni and the sputter deposited samples was determined

    by x-ray diffraction (XRD) using a Siemens D-500 Diffractometer with Cu K radiation. -2

    scans were performed from 20 to 80 with a step size of 0.008 and a hold time of 5 s. The

    approximate grain size was obtained using the Debye-Scherrer formula by measuring the full

    width at half maximum (FWHM) of the {111} peak in N-Ni, S-Ni, and T-Al and measuring the

    1010peak in titanium.The SMAT samples have a gradient microstructure, and therefore the obtained grain

    size from XRD represents the average size in the probed volume. The SMAT nickel and

    35

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    48/90

    aluminum were analyzed by comparing -2 scans of their {111} peak with a fixed 12 incident

    angle 2 scan of the same peak. The step size was 0.008 with a hold time of five seconds.

    The SMAT titanium was cut into quarters using an Allied Techcut 5 diamond saw, with a

    water-based coolant, at 800 RPM and a cutting speed of 0.5 cm/min. One quarter was then

    mounted in epoxy. Finally, the SMAT sample was sectioned with a diamond saw at the same

    conditions, at a slight angle with respect to the SMAT surface. The sectioning created a thin

    SMAT layer sample and a thin bulk layer from the back of the SMAT surface. Due to the angle

    of the cut, the thickness of the sample varied approximately from 250 m to 90 m. -2 scans

    were performed on both slices, with the SMAT slice positioned such that the x-rays penetrated

    near the 90 m edge. The grain size was approximated using a the Debye-Scherrer formula.

    4.3 Material Characterization

    4.3.1 Nanoindentation

    A Hysitron TI 700 Ubi1 nanomechanical test instrument was used to characterize the

    hardness of the samples. A Berkovich indenter tip was used to perform the measurements.

    The measurements characterized the hardness of the samples versus the depth from the

    surface. A single indent profile was used, and consists of a constant loading rate of 50 nm/s, a

    1 second hold, followed by a 25 nm/s unload. Figure 4.2 displays the loading profile graphically.

    The Hysitron software package, Tribosoft, analyzed the resulting load versus displacement

    data, and using a power law equation, calculated the hardness and reduced modulus. At least

    four indents were performed at any one indent depth and the average is reported.

    36

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    49/90

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    50/90

    are moving average trend line fits, using 100 COF data points. This is used to remove chatter

    of the graphs so that they can be easily read and compared.

    The wear volume for each test was calculated by averaging four 90-separated cross-

    sections of the wear track and multiplying it by the circumference of the track. The wear rate

    was then calculated by dividing the wear volume by the contact load and sliding distance. The

    wear rate of the alumina pin was calculated by measuring the wear scar diameter. Two

    measurements were taken, 90 apart and then averaged to obtain the scar diameter. Following

    ASTM G99-05, the wear pin volume was then calculated by equation 3.

    r

    d W P

    64

    4

    = (3)

    where WP = Wear Volume, d = average wear scar diameter, and r = wear pin radius.

    4.5 Hertzian Stress

    The initial stress on the sample surface from the loaded pin must be known to

    understand the wear test starting conditions. High Hertzian stresses can cause yielding of the

    surface material, which can affect the outcome of the test. The maximum contact stress

    between the pin and the surface is calculated by finding the maximum contact area between the

    pin and the surface. This is obtained by following a well established formula as presented by

    Shigley [56]. The radius of the contact area is calculated using equation 4.

    (4)F is the applied force, v is Poissons ratio, E is the modulus of elasticity, and d is the diameter of

    the contacting surfaces. The maximum contac tress then is calculated by equation 5.t s

    p F (5)

    The maximum deflection, maximum shear stress, and location of maximum shear stress can

    then be calculated as presented in Stachowiak and Batchelor [57].

    38

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    51/90

    The SMAT materials have a rough surface, and therefore the contact cannot be

    approximated as a sphere on a smooth surface. The SMAT surface topography is

    characterized by many dimples and peaks, or asperities. The maximum Hertzian contact stress

    occurs when the pin is positioned on one of these asperities. The Hertzian stress can then be

    modeled as a sphere on an infinitely long cylinder. Therefore, the Hertzian contact stress was

    calculated using the approximate contact stress calculations for elliptical contact areas

    developed by Hamrock and Dowson, and as described in Stachowiak and Batchelor [57].

    4.6 Scanning Electron Microscopy (SEM)

    A Hitachi S-3000N variable pressure scanning electron microscope equipped with

    energy-dispersive spectrometry (EDS) was utilized to analyze the wear track, wear debris, and

    quality of deposited films. An electron beam energy of 10 to 25 kV was used, depending on the

    material being analyzed. All samples were mounted on an aluminum stub with graphite tape.

    After the tests, wear debris was analyzed by placing the debris on copper tape, and then

    mounting the copper tape to an aluminum mount with graphite tape.

    4.7 Surface Profilometry

    The volume loss of the wear tracks was measured using a Veeco Wyco NT9100 optical

    profilometer. Each sample was scanned using the units VSI (Vertical Scanning interferometry)

    mode at various magnifications and scan lengths. The surface roughness of each sample was

    calculated by using the accompanying Veeco software, Vision. Four measurements were taken

    at 5.0X magnification, which corresponds to a surface area of 1.2 mm 2, and the results were

    averaged.

    39

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    52/90

    CHAPTER 5

    RESULTS AND DISCUSSION

    5.1 Microstructure

    The approximate grain size of the SMAT samples and the PVD aluminum and titanium

    samples was calculated by using x-ray diffraction. The results for the bulk, electrodeposited,

    and sputtered samples are displayed in Table 5.1 .

    S-Ni displays a XRD pattern similar to N-Ni, indicating that the sputtered film has a

    grain size in the nanoscale. S-Ti displays many oxide peaks along with the titanium peaks,

    which indicates that the sputtered titanium is possibly not fully dense and is oxidized. T-Al has

    a large grain size and an oxide layer is clearly present on the surface, along with XRD oxide

    peaks. This indicates that S-Ti and T-Al sputtered films are not high quality films.

    (c)(b)(a)

    Figure 5.1 Optical micrographs of etched (a) Ni-1 (b) Al-2 and (c) Ti-1.

    Table 5.1 Average grain size of Bulk, electrodeposited, and sputtered samples.

    Sample Grain SizeNi-1 39.2 3.1 mN-Ni 54.8 nmS-Ni 75.4 nm Al-2 291.3 30.7 mT-Al >200 nmTi-1 34.0 3.8 mS-Ti 126 nm

    [

    40

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    53/90

    Direct evidence of a gradient microstructure in the SMAT materials is difficult to obtain.

    Multiple studies indicate that SMAT produces a gradient microstructure [9-11, 22, 50-53]. XRD

    analysis can indicate that a gradient does exist. Peak broadening is a leading indication of

    grain refinement. A problem with regular -2 diffraction is the significant depth of penetration

    inhibits surface probing. With polycrystalline materials, the incident x-ray beam probes a large

    volume and Braggs law will always be satisfied due to the presence of many randomly oriented

    grains. Therefore, to avoid penetrating deep in the surface, a diffraction scan centered at 12

    with the surface was conducted for the SMAT nickel and SMAT aluminum samples. The results

    are shown in Figure 5.2 . Titanium has a higher x-ray absorption coefficient than nickel and

    aluminum, and a 12 scan provided only a weak peak. Therefore, the SMAT titanium was

    mounted in epoxy and sliced along the SMAT surface, creating a 100 m thin SMAT foil and a 1

    mm thick bulk foil. -2 scans were then performed on both slices, and the results were

    compared. This method provides an insight on the top layer of Ti-2 and the bulk titanium in Ti-

    2. Table 5.2 displays the results of the XRD analysis.

    (a) (b)

    Figure 5.2 -2 and low angle XRD results for (a) Ni-2 and (b) Al-3.

    41

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    54/90

    Figure 5.3 -2 scan of Ti-2 before (bulk SMAT Ti-2) and after (SMAT Layer Ti-2) the SMATsurface was removed.

    Table 5.2 Peak location, FWHM, and approximate grain size of SMAT samples from XRDanalysis.

    Scan Type 2

    (deg.) FWHM (deg.)

    Grain Size (nm)

    Ni-2{111}

    -2 Scan 44.43 0.25 34.9 = 12 44.39 0.34 25.2

    Al-3{111}

    -2 Scan 38.42 0.18 47.8 = 12 38.39 0.24 34.5

    Ti-2{ }

    -2 SMAT Edge 39.94 0.50 33.0 -2 Bulk 40.25 0.26 16.9

    The XRD scans on the SMAT nickel, aluminum, and titanium display both peak

    broadening and peak shifting. The peak shifting of the SMAT nickel and aluminum are similar in

    both scans, indicating that a residual stress is present in the material. This residual stress is

    likely from the SMAT processing. It is well known that residual stresses will shift XRD peak

    positions. A shift to a lower Bragg angle indicates a tensile stress is present. This at first

    seems unlikely, however, the SMAT samples were prepared in a vial, with the edges of the

    round sample fully clamped. As the SMAT process is applied, the material will deform, and

    since it is clamped, will cause the sample to bow out, creating a tensile stress near the SMAT

    surface and a compressive stress on the back surface. This can be seen in the SMAT titanium.

    The SMAT layer displays a shift lower than Braggs angle, and the back bulk part of Ti-2

    displays a shift higher than Braggs angle, indicating a compressive stress. This is not direct

    42

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    55/90

    evidence of a gradient microstructure, however, by the mechanics of the SMAT process, it is not

    likely that a thin layer of smaller grains forms at the surface, with no grain size gradient to form.

    5.2 Hardness

    5.2.1 Nanoindentation

    Since six indentations were performed at the same location but at different depths, work

    hardening is a concern. An existing SMAT nickel sample was used as a test specimen. The

    sample was processed for one hour with six 3 mm diameter harden steel balls. Eight individual

    indents were performed at each displacement of the sample, followed by eight indents using the

    multi-displacement loading curve. The results, displayed in Figure 5.4, show that there was

    only a small difference between the two tests. The difference in hardness can largely be

    attributed to the loading rate used for the tests. The individual indents were performed with a

    set loading, hold, and unloading time of five seconds, whereas the step indents were performed

    using the loading profile shown in Figure 4.2. There is a larger error with the step indents and a

    slight decrease in the hardness is observed. However, this is within error limits and therefore

    can be used.

    43

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    56/90

    (a)

    (b)

    Figure 5.4 Results of individual indentation tests compared to results of stepped tests on SMATNi; (a) Reduced modulus and (b) Hardness.

    44

  • 7/28/2019 Tribological Behavior of Nanocrystalline Metals Prepared by Surface Mechanical Xxx

    57/90

    The mc and SMAT materials were tested for hardness. The results are shown in Figure

    5.5 and Figure 5.6. The results show that the SMAT nickel hardness is lower than the mc-nickel

    and the SMAT aluminum and SMAT titanium are showing hardness values only slightly higher

    than their mc counterpart. When the reduced modulus is examined, the SMAT nickel modulus

    is drastically lower than the mc nickel, and deviations are seen in the SMAT aluminum and

    SMAT titanium. The change in reduced modulus is a sign that the rough SMAT surface may

    have affected the values